Application of molecular dynamics simulations to study

advertisement
Proceedings of the 2nd Russian-Bavarian Conference on Bio-Medical Engineering
Application of Molecular Dynamics Simulations to Study of
Membrane Related Structures
K.V. Shaitan, Ye.V. Tourleigh, A.K. Shaytan, K.B. Tereshkina, O.V. Levtsova, M.P.
Kirpichnikov
M.V. Lomonosov Moscow State University
Abstract: The paper gives an overview of approaches for applying molecular dynamics (MD)
simulations to study of membranes and biomembrane systems developed at the Department of
Bioengineering of the Biological Faculty. With reference to dynamic approaches to molecular
modeling and design of complex structures the following is discussed: biomembranes dynamics, direct
in silico free energy calculations. Diffusion phenomena in membrane structures are surveyed;
procedure of determination of aminoacid residues hydrophobicity at the phase boundary is discussed.
Methods for free energy estimates through molecular dynamics simulations are discussed.
Index Terms – CAMD, molecular dynamics, membranes, free energy calculation
Introduction
MD At present, MD and molecular modeling play an important role in the advancement of science and
improvement of technologies [1, 2]. The system containing a large number of classical equations of
motion for atomic particles is solved, as a rule, with the use of the Verlet difference scheme [3]. The
force field is set by a system of atom-atomic pair potentials, which are specially gauged for particular
type of molecular objects (biopolymers, minerals, alloys and so forth). Normally, special algorithms
for maintaining a constant temperature and pressure (or volume) are used in addition. By applying
special techniques free energy estimates can be performed with MD simulations [4]. In general, it is
rather relevant application of molecular modeling techniques for biotechnology, drug design etc. In
our case we can measure solvation free energies to determine the interaction properties of molecules in
various media, for instance with water or inner layers of membranes, and thus we can determine
hydrophobicity properties.
Along with the equilibrium MD, i.e. MD only under the effect of interatomic interactions, the
nonequilibrium (or steered) MD (SMD, [5]) under the action of additional forces and/or special
boundary conditions is applied. Usage of the nonequilibrium MD in application to complex and
heterogeneous molecular systems is now rather perspective. An overview of MD methods in
application to study of membrane molecular systems is given in the next sections.
Dynamics of Biomembranes
A. Membrane anisotropy and viscous properties
Biological membranes are a subject of rapt studying in recent years [6–10]. However, data on kinetic
characteristics of the phospholipid bilayer with account to its anisotropy and heterogeneity are quite
few. Below the problem of heterogeneity and anisotropy of a phospholipid bilayer is discussed with
respect to diffusion processes. A membrane consisting of 1-palmitoyl-2-oleoyl-sn-glycero-3phosphatidylcholine (POPC) bilayer is considered in details. The degree of solvation is 44 water
molecules per one lipid molecule. TIP3P water model is used; the valence bonds and valence angles in
water molecules are not fixed, but are instead determined by the corresponding potentials. The surface
density of lipids is close to the experimental values (62-68 Å2, [11–14]).
The molecular dynamics was calculated with the PUMA software package [15], which was specially
modified to include SMD [7, 8]. The system of classical equations for atom movement was solved in
the Amber 1999 force field [16]. Periodic boundary conditions, collisional thermostat [15] (T=300 K)
and Berendsen barostat [17] were applied. Fluctuations of volume, pressure and temperature were
used to verify that the system has reached a local equilibrium [7].
For computation of parameters which determine diffusion of molecules in the membrane, the SMD
method was used. External forces (constant and alternating) were applied to some parts of the system.
Trial spheres of 18 Da with radii of 2 and 4 Å (i.e., an order of the radius of the carbon atom and a
small functional group, respectively), which interact with the other atoms only by van der Waals
forces (interaction constant ε, 0.15 kcal/mol). A constant external force Fext of 0.3 to 4 kcal/molÅ-1
was applied to the spheres along the normal to the membrane plane (Fig. 1).
At molecular level essential deviations from the hydrodynamic Stokes formula are observed, and that
is quite natural, since the continuous medium approximation for the particles of such a size almost
does not hold. The Stokes equation can provide valid only qualitative estimates.
Also cases with a laterally applied force Fext of 1, 2, 4 and 10 kcal/molÅ-1 were studied. The
coefficient of viscous friction γ was defined as the ratio of external force to the velocity of particle
drift.
Formally, the friction coefficient could be redefined in terms of diffusion coefficient, using the known
Einstein relation, and in terms of microviscosity of the medium using the Stokes equation.
Currently, limited data are available on the viscosity when a particle moves along the normal to
membrane surface or in the lateral direction in the middle of the bilayer. Experimental averaged
viscosities of the surface layer range from 30 to 190 cP in various lipid membranes [18–20]. The
experimental estimate of the mean viscosity of POPC is about 18 cP [21]. Since the microviscosities
are not the same in different regions of the membrane, it is reasonable to define several structurally
and dynamically inhomogeneous regions. In the first approximation, one can distinguish the regions of
lipid headgroups and alkyl chains. Data on the friction coefficient in terms of microviscosity for
several parts of the system at various values of the external force are given in [7, 8].
Calculated values of effective viscosity of water for a 2-Å sphere are 0.3–0.4 cP, which is two to three
times lower than the experimental values. These data agree with the known estimates of water
viscosity for the TIP3P model [22]. Transverse viscosity of the membrane does not exceed 6 cP.
Viscosity of the central part of the bilayer is several times lower than this value.
Studying lateral movement of a sphere under the influence of a force results in the values of viscosity
which are close to those measured in the region of alkyl tails in the direction of the normal.
In general, the results suggest a non-Newtonian character of the medium and a weakly nonequilibrium
state of this system at movement velocities of 1–1-10 Å/ps.
The velocity of penetration of the molecule under the influence of external force also depends on the
chemical properties of the molecule. For comparison, the dynamics of penetration of tryptophan
(effective radius, 4.8 Å) and alanine residues (effective radius, 3.1 Å), as well as of the 2-Å van der
Waals sphere, into the bilayer was calculated.
It should be noted that a more polar tryptophan residue has a higher velocity in the region of lipid
headgroups than the alanine residue, which, consequently, provides a lower value of microviscosity. In
the region of hydrophobic alkyl chains, the situation is the reverse, with velocities differing by a factor
of 15. The region of lipid headgroups is most sensitive to the nature of a molecule moving through the
membrane. Conversely, the hydrophobic core of the bilayer with a larger free volume is sensitive to
the size of particles.
B. Lateral self-diffusion of lipids in bilayer
Far more detailed experimental information on lateral diffusion coefficients of lipids in membranes is
available. These data can be used for testing an MD technique. Lateral self-diffusion coefficient Dxy of
lipids is determined according to the known relation
 ( x(t )  x(0))2  ( y(t )  y(0))2  4 Dxyt ,
Proceedings of the 2nd Russian-Bavarian Conference on Bio-Medical Engineering
where mean squared deviation of lipids’ centers of mass in the plane of the bilayer is in angle brackets
The averaging is done for all the lipids. A trajectory 3 ns long is divided into 12 parts 250 ps long
each. Dependence of squared displacement (Fig. 2) was obtained after averaging over all 12 parts.
The value of Dxy = 2.4ּ10-7 cm2/s calculated for POPC is close to that from the data on quasi-elastic
neutron
scattering
on
dipalmitoylphosphatidylcholine
(1×10-7
см2/с
[23])
and
-7
2
dioleoylphosphatidylcholine (2×10 cm /s [24]) bilayers. The results of pulsed NMR for POPC
bilayer provide with values of 2.0×10-7 cm2/s at 298 K and 2.5×10-7 cm2/s at 303 K [25], which
practically coincide with the value obtained in the simulation. Let us note that the obtained value of
lateral diffusion coefficient turned out to be closer to the experimental values than the coefficients out
from MD simulations conducted according to other techniques (3±0.6×10-7 cm2/s [26], 4.0-4.5ּ10-7
cm2/s [27] for dipalmitoylphosphatidylcholine bilayers and 7.3ּ10-5 cm2/s [10] for POPC bilayer).
Estimation of Aminoacid Residues Hydrophobic Properties at the
Water-Membrane Interface
Let us consider an example of technique of special boundary conditions. The behavior of single
aminoacid residues at the interface between water and a non-polar solvent was studied by means of
SMD.
Simulation was carried out both for the water-hexane interface and for a model system water-vacuum.
In the latter case the aqueous layer was separated from vacuum by a flat virtual wall serving as an
additional repulsive potential for water molecules. At that, the wall remained permeable for aminoacid
residues. Collisional thermostat was applied to molecules, including those in vacuum. Thus, the
absence of hydrophobic environment was partly compensated.
Aminoacid residues were taken in the form of biradical monomer unit -NH-CH (R)-CO-, that is as
uncharged monomers with formally unsaturated bonds. PUMA software was used. Periodic boundary
conditions, Amber 1999 force field and TIP3P water model with non-fixed internal degrees of
freedom were chosen. Berendsen barostat was used along the axis perpendicular to the phase boundary
in the water-hexane system. The water-vacuum system was simulated in NVT-ensemble, the density of
water corresponded to standard conditions.
For estimation of hydrophobic properties of aminoacid residues, statistical handling of trajectories was
done and space and orientation distributions of residues were analyzed. A comparison of these two
models follows below.
Let us consider the behavior of a phenylalanine molecule at the water-hexane phase boundary (Fig. 3).
Distributions of the centers of mass of the phenylalanine’s backbone and side chain are shown on Fig.
4. It is seen that the phase boundary represents a transition layer about 5 Å thick. Maxima of the
distributions practically correspond to the middle of the interfacial layer, that is the molecule reveals
surface-active properties. The distribution plots for the hydrophobic side chain and the polar backbone
are shifted towards the hexane and the aqueous phase, respectively. Analysis of molecular orientations
shows that the side chain is predominantly directed towards the hexane. The vector connecting the
centers of mass of the backbone and the side chain is at an angle of 30 degrees with respect to the
phase boundary, as if the molecule almost laid on the interface.
Let us proceed with the dynamics of the phenylalanine in the system with a virtual repulsive wall. For
such a system, the steepness of the water phase density profile is more than in the water-hexane
system. However, the absence of non-polar hexane phase insignificantly changes the shape and width
of the aminoacid residues distributions. The most probable positions of the centers of mass of the
backbone and the side radical are shifted towards the aqueous phase in comparison with the case of the
water-hexane interface. Thus, the most probable position of the molecule is in the water layer adjacent
to the interface but not at the interface itself. The dynamics of molecular orientation in the watervacuum and water-hexane systems differs insignificantly.
The model of virtual repulsive plane can be used for fast estimation of hydrophobic and surface-active
properties of molecular structures. This model was also used for analysis of properties and
classification of hydrophobic properties of the majority of aminoacid residues. The estimation of
hydrophobic parameters is used for plotting hydrophobicity profiles of proteins [28, 29] as well as it is
applied to pharmacokinetics [30]. Hydrophobicity is one of the major parameters in reactions
catalyzed by enzymes. Binding sites of the latter are rather sensitive to hydrophobic areas of substrate
molecules [31]. While analyzing the distributions, it was found that residues of all main types of
amino acids can be split, to a first approximation, into two groups. The first one reveals evident
surface-active properties. All residues with non-polar side chains, including glycine, belong to this
group. The residues of the second group desorbe from the surface and leave to the aqueous phase. This
group is comprised of the residues, the side groups of which are either polar or charged. We underline
that the developed approach allows developing a system of quantitative indicators characterizing
hydrophobic properties of molecules.
Free Energy Calculations for Molecular Modeling
Theoretical description of processes taking place in many-particle (thermodynamic) systems is based
on the concepts of entropy and free energy. In fact, a certain value of free energy can be attributed to
every state of a thermodynamical system so that it would be somewhat equivalent to the probability of
finding the system in this particular state. Free energy is widely used as a quantitative measure of
affinity between substances. For instance, free energy of solvation for a given solvent in a given solute
determines the affinity of solute to this or that liquid. Imagine we put to immiscible liquids in a contact
and solvate small amount of substance in this biphasic system, then, provided we know solvation free
energies for each solvent/solute pair, we can determine the concentration of this substance in each
liquid. Generally, the dissolved substance will tend to concentrate in that phase, where its solvation
free energy is lower. Determination of hydration free energy will give us information about the
interaction of the substance with water at the macroscopic level; this can be used for design of new
hydrophobic materials and surfaces. The derived interaction properties can be further used in coarsegrained models for prediction of behavior of more complex compounds, such as polymeric systems,
and design of smart polymers.
Analogous processes, but much more complicated ones, take place in biological systems. Of utmost
importance for biomedical applications is determination of binding affinity between ligands and
proteins/receptors which is actually the basis of rational drug design [32]. Drug design is often reduced
to finding a ligand that binds with the specific protein site. The binding strength is actually the binding
free energy. Provided we know binding free energies for different substances, then we can decide what
substance would perform better.
All the above stated examples suggest that a technique enabling us to determine precise free energy
estimations just in silico, by means of computer simulations, would greatly facilitate design of
technologically and biologically relevant compounds as well as serve a good means of scientific
investigation.
Today the computational power has developed to such extent that we can speak about determining the
free energy properties from molecular dynamics simulations with statistical error compared to
experimental one.
In general precise free energy estimates require significant computational power. But however it
should be stated that the choice of algorithm for free energy estimates in MD simulations as well as
the choice of force field and other simulation parameters can considerably affect accuracy of estimates
and the amount of required computational resources. Study of possible methods, algorithms for free
energy estimates through computer simulations and their efficacy with respect to precision/cost ratio is
performed in our laboratory.
In this article we would not dwell much on the theory of free energy estimations and comparison of
different methods, but rather consider a simple example just to give the reader a notion how the
technique looks like and what are the computational costs. The example deals with determination of
hydration free energy of a simple methane molecule.
At first step we should mention that we deal with classical MD simulations, it means that a molecular
system is represented by a number of atoms (point particles) that interact with each other according to
predefined potential energy functions. Definition of these energies and hence forces is taken from
Proceedings of the 2nd Russian-Bavarian Conference on Bio-Medical Engineering
chosen force field parameters. For our simulation we have chosen the OPLS-AA parameter force field
for treating methane molecule and water of SPC type.
From the theory we know that the problem of determining hydration free energy can be reduced to
finding the free energy difference between a system of one methane molecule solvated in some
amount of water and the same system where the methane molecule does not interact with water
molecules at all.
So the next step is to construct this system in silico and topologies (force field parameters) for its two
variations.
We will now use the “thermodynamic integration” (TI) method for determining the free energy
difference. The brief theory of the method is as follows. From the statistical physics we know that we
can estimate free energies F in molecular modeling using formula (1), where the integral is taken over
all the phase space of the system and Hamiltonian is just the energy of the system in each particular
state. Theoretically this integral can be evaluated numerically for small systems, but practically it is
almost impossible.
F   kT ln  e

Hamiltonian
kT
dpdq
(1)
One of the methods that help to overcome this problem is the TI method. The practical implementation
of the method should consist of the following steps. Suppose we would like to find the free energy
differences of two different systems at equilibrium. We should construct a special parameter (λ)
dependent Hamiltonian for our system, so that at one value of λ it refers to the first system and at
another value in refers to the second system. Than we should make multiple MD simulation runs with
Hamiltonian referring to different values of λ. During each MD run we should sample the partial
derivative of Hamiltonian with respect to λ and in the end integrate the sampled function over the
same parameter (2)
1
F  
0
H
d (2)

An important conclusion from this consideration is that those MD simulation runs can be made in
parallel. In fact the estimation of any statistical property from MD simulation being an average over a
trajectory can be split into to separate independent runs and than averaged over these runs. This makes
it possible to apply parallel computing at this point.
The next important step is the choice of simulation parameters. The most important of them include
treatment of non-bonded interactions (van der Waals and Coulomb terms) and temperature coupling.
Using “cut-off” treatment for non-bonded interactions from one side can decrease simulation time but
from the other introduces a systematic error in our measurements neglecting the “tails” of interaction.
That is why highly preferable is the use of special methods such as Ewald summation, dispersion
correction and others that help to treat long range interactions correctly. It should be also noted that the
choice of parameterization of Hamiltonian with respect to λ is also quite important. The right choice
can save computational power and improve accuracy.
After all the preparation work is done, we can run simulations in parallel for different values of λ. We
use for this purpose a “beowulf” type computer cluster. It should be also noted that after
measurements are done not only the value of sampled property should be estimated but its statistical
and systematic uncertainty as well.
The result of the above described process performed mainly with GROMACS software package [33]
is depicted on Fig. 5. The result of the simulation (after integrating the curve) is the following. The
estimate for hydration free energy of methane in water, using OPLS-AA/SPC forcefield in our
simulation is 2.20 kcal/mol with statistical standard deviation of 0.02 kcal/mol. The total amount of
computational resources used for this estimate is approximately 30 day of processor time, processor
type Intel Xeon 3.2 GHz.
Comparing with experimental value of 1.94 kcal/mol [34], we see that our result overestimates the
hydration free energy a bit. The reason apparently lies in the choice of force field. For
parameterization of the majority of cotemporary force field data on hydration free energies was not
used, because when they were elaborated the computational resources were not enough to determine
free energies with high accuracy. Now, with development of computer sciences and equipment it is
becoming possible, and we can not only pose the question of free energy estimation in computer
simulations but also the problem of elaboration of new force fields and parameter sets for molecular
substances on the basis of this knowledge.
Conclusion
Molecular dynamics and its different variations (nonequilibrium (or steered) MD, thermodynamic
integration) appears to be a useful tool for obtaining detailed and experimentally hardly accessible
information on structure, dynamic, kinetic, statistical properties of heterogeneous and anisotropic
systems such as biomembranes.
Available data show that the given method appears quite useful for prognosis of structural and
functional properties of a broad class of molecular systems and can be used for molecular design of
objects on the nanometer scale.
Acknowledgement
The authors thank the Russian Foundation for Basic Research (grant № 04-04-49645), RF Ministry of
Education and Science, Rosnauka federal agency, CRDF (USA) and Moscow Government for the
financial support.
References
1.
H. Gao, “Modelling strategies for nano- and biomaterials”, in European white book on
fundamental research in materials science, Van der Woorde M.H. et al., Eds. Max Planck
Gesellschaft, 2001, pp.144–148.
2.
K.V. Shaitan, Ye.V. Tourleigh, D.N. Golik, K.B. Tereshkina, O.V. Levtsova, I.V. Fedik, A.K.
Shaytan, A. Li, and M.P. Kirpichnikov, “Dynamis and molecular design of bio- and nanostructures”,
Rossiyskiy Khimicheskiy Zhurnal (in Russian), vol. 50. pp. 53–65, April 2006.
3.
D.Frenkel and B. Smit, Understanding molecular simulation: from algorithms to applications,
2nd ed. San Diego: Academic Press, 2002.
4.
K.V. Shaitan, and K.B. Tereshkina, Molekulyarnaya dinamika belkov i peptidov, Мoscow:
Oykos, 2004.
5.
S. Park, and K. Schulten, “Calculating potentials of mean force from steered molecular
dynamics simulations”, J. Chem. Phys., vol. 120, pp. 5946–5961, April 2004.
6.
Y.A. Chizmadzhev, “The mechanisms of lipid-protein rearrangements during viral infection”,
Bioelectrochem., vol. 63, pp. 129–136, June 2004.
7.
Ye.V. Tourleigh, K.V. Shaitan, and N.K. Balabaev, “Molecular dynamics of a hydrated
palmitoyloleoylphosphatidylcholine bilayer in collisional thermostat”, Biologicheskie Membrany (in
Russian), vol. 22, pp. 491–502, December 2005.
8.
Ye.V. Tourleigh, K.V. Shaitan, and N.K. Balabaev, “Dynamic heterogeneity of phospholipid
bilayer and diffusion of molecules at the interface”, Biophysics, vol. 50, pp. 899–904, December 2005.
9.
A.L. Rabinovich, N.K. Balabaev, M.G. Alinchenko, V.P. Voloshin, N.N. Medvedev, and P.
Jedlovszky, “Computer simulation study of intermolecular voids in unsaturated phosphatidylcholine
lipid bilayers”, J. Chem. Phys., vol. 122, pp. 84906, February 2005.
10. H. Heller M., Schaefer, and K. Schulten, “Molecular dynamics simulation of a bilayer of 200
lipids in the gel and in the liquid crystal phases”, J. Phys. Chem., vol. 97, pp. 8343–8360, August
1993.
11. G. Pabst, M. Rappolt, H. Amenitsch, and P. Laggner, “Structural information from multilamellar
liposomes at full hydration: full q-range fitting with high quality X-ray data”, Phys. Rev. E., vol. 62,
pp. 4000–4009, April 2000.
Proceedings of the 2nd Russian-Bavarian Conference on Bio-Medical Engineering
12. B. Isralewitz, M. Gao, and K. Schulten, “Steered molecular dynamics and mechanical functions
of proteins”, Curr.Opin. Struct. Biol., vol. 11, pp. 224–230, April 2001.
13. J.M. Smaby, M.M. Mornsen, H.L. Brockman, and R.E. Brown, “Phosphatidylcholine acyl
unsaturation modulates the decrease in interfacial elasticity induced by cholesterol”, Biophys. J., vol.
73, pp.1492–1505, September 1997.
14. P.A. Hyslop, B. Morel, and R.D. Sauerheber, “Organization and interaction of cholesterol and
phosphatidylcholine in model bilayer membrane”, Biochemistry, vol. 29, pp. 1025–1038, January
1990.
15. A.S. Lemak and N.K. Balabaev, “A comparison between collisional dynamics and Brownian
dynamics”, Mol. Simul., vol. 15, pp. 223–231, March 1995.
16. J. Wang, P. Cieplak, and P.A. Kollman, “How well does a restrained electrostatic potential
(RESP) model perform in calculating conformational energies of organic and biological molecules?”,
J. Comp. Chem., vol. 21, pp. 1049–1074, July 2000.
17. H.J.C. Berendsen, J.P.M. Postma, W.F. van Gunsteren, A. DiNola, and J.R. Haak, “Molecular
dynamics with coupling to an external bath”, J. Chem. Phys., vol. 81, pp. 3684–3690, October 1984.
18. J.F Nagle, R. Zhang, S. Tristram-Nagle, W.J. Sun, H.I. Petrache, and R.M. Suter, “X-ray
structure determination of fully hydrated Lα phase dipalmitoylphosphatidylcholine bilayers”, Biophys.
J., vol. 70, pp. 1419–1431, March 1996.
19. M.C. Wiener, and S.H. White, “Structure of a fluid dioleoylphosphatidylcholine bilayer
determined by joint refinement of X-ray and neutron diffraction data. III. Complete structure”,
Biophys. J., vol. 61, pp. 434–447, February 1992.
20. G. Reiter, M. Siam, D. Falkenhagen, W. Gollneritsch, D. Baurecht, and U.P. Fringeli,
“Interaction of a bacterial endotoxin with different surfaces investigated by in situ fourier transform
infrared attenuated total reflection spectroscopy”, Langmuir, vol. 18, pp. 5761–5771, June 2002.
21. A. Seelig and J. Seelig, “Effect of a single cis double bond on the structures of a phospholipid
bilayer”, Biochemistry, vol. 16, pp. 45–50, January 1977.
22. J. Seelig and A. Seelig, “Lipid conformation in model membranes and biological membranes”,
Q. Rev. Biophys., vol. 13, pp. 19–61, February 1980.
23. E. Sackmann, “Structure and Dynamics of Membranes”, in Handbook of Biological Physics,
vol. 1, R. Lipowsky and E. Sackmann, Eds. Amsterdam: Elsevier Sci. Pubs. B.V., 1995.
24. W. Pfeiffer, T. Henkel, E. Sackmann, W. Knoll, and D. Richter, “Local dynamics of lipid
bilayers studied by incoherent quasi-elastic neutron scattering”, Europhys. Lett., vol. 8, pp. 201–206,
October 1989.
25. A. Filippov, G. Oradd, and G. Lindblom, “The effect of cholesterol on the lateral diffusion of
phospholipids in oriented bilayers”, Biophys. J., vol. 84, pp. 3079–3086, May 2003.
26. U. Essmann and M.L. Berkowitz, “Dynamical properties of phospholipid bilayers from
computer simulation”, Biophys. J., vol. 76, pp. 2081–2089, April 1999.
27. T. Róg, K. Murzyn, and M. Pasenkiewicz-Gierula, “Molecular dynamics simulations of charged
and neutral lipid bilayers: treatment of electrostatic interactions”, Acta Biochim. Polon., vol. 53, pp.
789–798, September 2003.
28. J. Kyte and R.F. Doolittle, “A simple method for displaying the hydrophobic character of a
protein”, J. Mol. Biol., vol. 157, pp. 105–142, May 1982.
29. T.P. Hopp and K.R. Woods, “Prediction of protein antigenic determinants from amino acid
sequences”, Proc. Nat. Acad. Sci. USA, vol. 78, pp. 3824–3828, June 1981.
30. T. Sotomatsu-Niwa and A. Ogino, “Multiple regression analysis of the beta-sheet propensity of
amino acids”, J. Mol. Struct. (Theochem), vol. 392, pp. 43–54, December 1997.
31. L. Stryer, Biochemistry, 4th ed. New York: W.H. Freeman & Co., 1995.
32. M.R. Reddy and M.D. Erion, Free energy calculations in rational drug design. New York:
Kluwer Academic/Plenum, 2001.
33. D. van der Spoel, E. Lindahl, B. Hess, G. Groenhof, A.E. Mark, and H.J.C. Berendsen,
“GROMACS: fast, flexible and free”, J. Comp. Chem., vol. 26, pp. 1701–1718, December 2005.
34. R. Wolfenden, L. Andersson, P.M. Cullis, C.C.B. Southgate, “Affinities of amino acid side
chains for solvent water”, Biochemistry, vol. 20, pp. 849–855, February 1981.
a
b
c
Fig. 1
Fig. 2
Proceedings of the 2nd Russian-Bavarian Conference on Bio-Medical Engineering
Fig. 3
Fig. 4
Fig. 5
FIGURE CAPTIONS
Fig. 1. (a) Structure of POPC bilayer.(b) Kinetics of movement of the trial van der Waals 2-Å sphere,
subjected to a force of 10 kcal/molÅ-1 applied in the direction of the normal z. The center of the
bilayer is located at z = 0; the borders, at z =±20 Å. (b) The velocity of sphere movement averaged
over an interval of 0.1 ps.
Fig. 2. Mean squared displacement of lipid’s center of mass in the plane of the POPC bilayer. The
straight line is a linear approximation.
Fig. 3. Density profiles of water and hexane along the axis perpendicular to the phase boundary. The
density is normalized to unity.
Fig. 4. Probability density for position of center of mass of aminoacid residue’s backbone and side
chain in the water-hexane system.
Fig. 5. The resulting plot of TI method. We obtain the averaged partial derivative of Hamiltonian with
respect to λ at different values of λ. Statistical uncertainties are depicted as well.
Download